Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 19 de 19
Filter
Add more filters










Publication year range
1.
Food Microbiol ; 66: 40-47, 2017 Sep.
Article in English | MEDLINE | ID: mdl-28576371

ABSTRACT

Three selective enrichment methods, the United States Food and Drug Administration's (FDA method), the United States Department of Agriculture Food Safety Inspection Service's (USDA method), and the EN ISO 11290-1 standard method, were assessed for their suitability for recovery of Listeria monocytogenes from spiked mung bean sprouts. Three parameters were evaluated; the enrichment L. monocytogenes population from singly-spiked sprouts, the enrichment L. monocytogenes population from doubly-spiked (L. monocytogenes and Listeria innocua) sprouts, and the population differential resulting from the enrichment of doubly-spiked sprouts. Considerable L. monocytogenes inter-strain variation was observed. The mean enrichment L. monocytogenes populations for singly-spiked sprouts were 6.1 ± 1.2, 4.9 ± 1.2, and 6.9 ± 2.3 log CFU/mL for the FDA, USDA, and EN ISO 11290-1 methods, respectively. The mean L. monocytogenes populations for doubly-spiked sprouts were 4.7 ± 1.1, 5.5 ± 1.3, and 4.6 ± 1.4 log CFU/mL for the FDA, USDA, and ISO 11290-1 enrichment methods, respectively. The corresponding mean population differentials were 2.8 ± 1.1, 3.3 ± 1.3, and 3.6 ± 1.4 Δlog CFU/mL for the same three enrichment methods, respectively. The presence of L. innocua and resident microorganisms on the sprouts negatively impacted final levels of L. monocytogenes with all three enrichment methods.


Subject(s)
Food Contamination/analysis , Hazard Analysis and Critical Control Points/methods , Listeria monocytogenes/isolation & purification , Vegetables/microbiology , Vigna/microbiology , Consumer Product Safety/legislation & jurisprudence , Consumer Product Safety/standards , Listeria monocytogenes/classification , Listeria monocytogenes/genetics , Seeds/growth & development , Seeds/microbiology , United States , United States Department of Agriculture , United States Food and Drug Administration , Vigna/growth & development
2.
J Food Prot ; 79(11): 1904-1910, 2016 11.
Article in English | MEDLINE | ID: mdl-28221922

ABSTRACT

Microbial competition during selective enrichment negatively affects Listeria monocytogenes populations and may hinder the subsequent detection or recovery of this organism. Competition assays among 10 selected strains of Listeria and Citrobacter braakii were performed in buffered Listeria enrichment broth, 3-(N-morpholino)propanesulfonic acid-buffered Listeria enrichment broth, University of Vermont medium-modified Listeria enrichment broth, and Fraser broth. The individual contributions of each selective agent in these media were also assessed, as well as the contribution of incubation temperature. Acriflavine hydrochloride and sodium nalidixate were ineffective at preventing the overgrowth of C. braakii ; this resulted in substantially lower populations of Listeria than when the competitor was absent. At the higher levels, both of these selective agents were detrimental to Listeria populations. The highest enrichment populations of Listeria were observed when either NaCl or LiCl was present. In the absence of selective agents, the final populations of Listeria following competitive growth with C. braakii were not substantially affected by temperature; however, in the presence of selective agents, the Listeria populations were statistically higher at the higher incubation temperature. There are a limited number of selective agents available for use in Listeria -specific enrichment media, resulting in formulations that are only somewhat selective for this species. The optimization of current formulations may help researchers to improve Listeria recovery, particularly from products with a high microbial load. The understanding of the behavior and interactions between target and nontarget microorganisms in the presence of these available selective agents is a necessary step in the optimization of Listeria selective enrichment formulations.


Subject(s)
Food Microbiology , Listeria , Acids , Colony Count, Microbial , Culture Media , Listeria monocytogenes
3.
Food Microbiol ; 46: 528-534, 2015 Apr.
Article in English | MEDLINE | ID: mdl-25475325

ABSTRACT

The presence of multiple species of Listeria in regulated food products is not uncommon and can complicate the recovery of Listeria monocytogenes particularly on a non-differentiating medium. The potential complications of Listeria seeligeri and Listeria welshimeri on the recovery of L. monocytogenes from inoculated food test samples using the U.S. Food and Drug Administration's (FDA) selective enrichment procedure was investigated. Post-enrichment enumeration, in the absence of food product, indicates that some L. seeligeri and L. monocytogenes pairings may have population differentials as great as 2.7 ± 0.1 logs with L. seeligeri being the predominant species. A similar observation was noted for L. welshimeri and L. monocytogenes pairings which resulted in population differentials as large as 3.7 ± 0.2 logs with L. welshimeri being the predominant species. Select strain pairings were used to inoculate guacamole, crab meat, broccoli, and cheese with subsequent recovery by the FDA Bacteriological Analytical Manual (BAM) method with 10 colonies per sample selected for confirmation. The presence of L. seeligeri had little effect on the recovery of L. monocytogenes. The presence of L. welshimeri resulted in the failure to recover L. monocytogenes in three out of the four food matrices. This work extends the observation that non-pathogenic species of Listeria can complicate the recovery of L. monocytogenes and that competition during selective enrichment is not limited to the presence of just Listeria innocua.


Subject(s)
Food Contamination/analysis , Food Microbiology , Listeria monocytogenes/isolation & purification , Listeria/growth & development , Listeria monocytogenes/growth & development
4.
Food Microbiol ; 44: 173-9, 2014 Dec.
Article in English | MEDLINE | ID: mdl-25084660

ABSTRACT

The growth of Listeria monocytogenes during the pathogen specific enrichment of food samples can be limited by the presence of additional microorganisms that are resistant to the selective conditions being applied. If growth is severely limited and minimum post-enrichment threshold levels are not met then the presence of L. monocytogenes may go undetected. Several food products were screened for non-pathogenic commensal or spoilage microorganisms that are capable of growth under the conditions commonly used by regulatory testing laboratories to select for Listeria species. The effect of these potential competitor microorganisms on the ability to detect L. monocytogenes by several common molecular screening assays was then determined. Eight species of bacteria were isolated from foods that demonstrated the ability to grow in buffered Listeria enrichment broth under selective conditions. Growth of these competitor microorganisms during the enrichment incubation resulted in a decrease ranging from 1 to 4 logs in the 48 h population of L. monocytogenes. Three strains of L. monocytogenes representing serotypes 1/2a, 1/2b, and 4b were included in this study but no one serotype appeared to be most or least sensitive to the presence of competitor microorganisms. One additional strain of L. monocytogenes was identified as displaying minimal growth during the enrichment period in the presence of the Citrobacter braakii with the final population only reaching approximately 2.6 log CFU/ml after 48 h which was a 2 log increase over the initial population. This particular strain was subsequently shown to be difficult to detect following enrichment by an automated immunofluorescence assay and an antibody-based lateral flow device assay. In some enrichments, this strain was also difficult to detect by real-time PCR.


Subject(s)
Bacteria/growth & development , Culture Media/metabolism , Listeria monocytogenes/growth & development , Bacteria/classification , Bacteria/isolation & purification , Bacteria/metabolism , Food Contamination/analysis , Listeria monocytogenes/metabolism
5.
J Food Prot ; 76(11): 1854-62, 2013 Nov.
Article in English | MEDLINE | ID: mdl-24215687

ABSTRACT

The recovery of low levels of Listeria monocytogenes from foods is complicated by the presence of competing microorganisms. Nonpathogenic species of Listeria pose a particular problem because variation in growth rate during the enrichment step can produce more colonies of these nontarget cells on selective and/or differential media, resulting in a preferential recovery of nonpathogens, especially Listeria innocua. To gauge the extent of this statistical barrier to pathogen recovery, 10 isolates each of L. monocytogenes and L. innocua were propagated together from approximately equal initial levels using the current U. S. Food and Drug Administration's enrichment procedure. In the 100 isolate pairs, an average 1.3-log decrease was found in the 48-h enrichment L. monocytogenes population when L. innocua was present. In 98 of the 100 isolate pairs, L. innocua reached higher levels at 48 h than did L. monocytogenes, with a difference of 0.2 to 2.4 log CFU/ml. The significance of these population differences was apparent by an increase in the difficulty of isolating L. monocytogenes by the streak plating method. L. monocytogenes went completely undetected in 18 of 30 enrichment cultures even after colony isolation was attempted on Oxoid chromogenic Listeria agar. This finding suggests that although both Listeria species were present on the plate, the population differential between them restricted L. monocytogenes to areas of the plate with confluent growth and that isolated individual colonies were only L. innocua.


Subject(s)
Culture Media , Food Contamination/analysis , Food Microbiology , Listeria monocytogenes/isolation & purification , Agar , Buffers , Colony Count, Microbial , Consumer Product Safety , Listeria/isolation & purification , Listeria monocytogenes/growth & development , United States
6.
Food Microbiol ; 36(2): 231-40, 2013 Dec.
Article in English | MEDLINE | ID: mdl-24010602

ABSTRACT

Use of 16S rRNA partial gene sequencing within the regulatory workflow could greatly reduce the time and labor needed for confirmation and subtyping of Listeria monocytogenes. The goal of this study was to build a 16S rRNA partial gene reference library for Listeria spp. and investigate the potential for 16S rRNA molecular subtyping. A total of 86 isolates of Listeria representing L. innocua, L. seeligeri, L. welshimeri, and L. monocytogenes were obtained for use in building the custom library. Seven non-Listeria species and three additional strains of Listeria were obtained for use in exclusivity and food spiking tests. Isolates were sequenced for the partial 16S rRNA gene using the MicroSeq ID 500 Bacterial Identification Kit (Applied Biosystems). High-quality sequences were obtained for 84 of the custom library isolates and 23 unique 16S sequence types were discovered for use in molecular subtyping. All of the exclusivity strains were negative for Listeria and the three Listeria strains used in food spiking were consistently recovered and correctly identified at the species level. The spiking results also allowed for differentiation beyond the species level, as 87% of replicates for one strain and 100% of replicates for the other two strains consistently matched the same 16S type.


Subject(s)
Bacterial Typing Techniques/methods , DNA, Bacterial/genetics , Listeria/isolation & purification , RNA, Ribosomal, 16S/genetics , Cheese/microbiology , Food Contamination/analysis , Fruit/microbiology , Listeria/classification , Listeria/genetics , Molecular Sequence Data , Phylogeny , Seafood/microbiology , Sequence Analysis, DNA
7.
Appl Microbiol Biotechnol ; 97(8): 3677-86, 2013 Apr.
Article in English | MEDLINE | ID: mdl-23494620

ABSTRACT

Alternative ligands such as nucleic acid aptamers can be used for pathogen capture and detection and offer advantages over antibodies, including reduced cost, ease of production and modification, and improved stability. DNA aptamers demonstrating binding specificity to Salmonella enterica serovar Typhimurium were identified by whole-cell-systematic evolution of ligands by exponential enrichment (SELEX) beginning with a combinatorial library of biotin-labeled single stranded DNA molecules. Aptamer specificity was achieved using whole-cell counter-SELEX against select non-Salmonella genera. Aptamers binding to Salmonella were sorted, cloned, sequenced, and characterized for binding efficiency. Out of 18 candidate aptamers screened, aptamer S8-7 showed relatively high binding affinity with an apparent dissociation constant (K d value) of 1.73 ± 0.54 µM and was selected for further characterization. Binding exclusivity analysis of S8-7 showed low apparent cross-reactivity with other foodborne bacteria including Escherichia coli O157: H7 and Citrobacter braakii and moderate cross-reactivity with Bacillus cereus. Aptamer S8-7 was successfully used as a ligand for magnetic capture of serially diluted Salmonella Typhimurium cells, followed by downstream detection using qPCR. The lower limit of detection of the aptamer magnetic capture-qPCR assay was 10(2)-10(3) CFU equivalents of Salmonella Typhimurium in a 290-µl sample volume. Mean capture efficiency ranged from 3.6 to 12.6 %. Unique aspects of the study included (a) the use of SELEX targeting whole cells; (b) the application of flow cytometry for aptamer pool selection, thereby favoring purification of ligands with both high binding affinity and targeting abundant cell surface moieties; and (c) the use of pre-labeled primers that circumvented the need for post-selection ligand labeling. Taken together, this study provides proof-of-concept that biotinylated aptamers selected by whole-cell SELEX can be used in a qPCR-based capture-detection platform for Salmonella Typhimurium.


Subject(s)
Aptamers, Nucleotide , Bacteriological Techniques/methods , Flow Cytometry/methods , Salmonella typhimurium/isolation & purification , Cross Reactions , Food Microbiology/methods , Sensitivity and Specificity
8.
Appl Microbiol Biotechnol ; 87(6): 2323-34, 2010 Aug.
Article in English | MEDLINE | ID: mdl-20582587

ABSTRACT

The need for pre-analytical sample processing prior to the application of rapid molecular-based detection of pathogens in food and environmental samples is well established. Although immunocapture has been applied in this regard, alternative ligands such as nucleic acid aptamers have advantages over antibodies such as low cost, ease of production and modification, and comparable stability. To identify DNA aptamers demonstrating binding specificity to Campylobacter jejuni cells, a whole-cell Systemic Evolution of Ligands by EXponential enrichment (SELEX) method was applied to a combinatorial library of FAM-labeled single-stranded DNA molecules. FAM-labeled aptamer sequences with high binding affinity to C. jejuni A9a as determined by flow cytometric analysis were identified. Aptamer ONS-23, which showed particularly high binding affinity in preliminary studies, was chosen for further characterization. This aptamer displayed a dissociation constant (K(d) value) of 292.8 +/- 53.1 nM with 47.27 +/- 5.58% cells fluorescent (bound) in a 1.48-microM aptamer solution. Binding assays to assess the specificity of aptamer ONS-23 showed high binding affinity (25-36%) for all other C. jejuni strains screened (inclusivity) and low apparent binding affinity (1-5%) with non-C. jejuni strains (exclusivity). Whole-cell SELEX is a promising technique to design aptamer-based molecular probes for microbial pathogens without tedious isolation and purification of complex markers or targets.


Subject(s)
Aptamers, Nucleotide/chemistry , Campylobacter jejuni/chemistry , Campylobacter jejuni/isolation & purification , SELEX Aptamer Technique/methods , Animals , Aptamers, Nucleotide/genetics , Campylobacter jejuni/genetics , Kinetics , Nucleic Acid Conformation , Poultry/microbiology , Species Specificity
9.
Proc Natl Acad Sci U S A ; 104(12): 4840-5, 2007 Mar 20.
Article in English | MEDLINE | ID: mdl-17360343

ABSTRACT

Type II DNA topoisomerases are essential and ubiquitous enzymes that perform important functions in chromosome condensation and segregation and in regulating intracellular DNA supercoiling. Topoisomerases carry out these DNA transactions by passing one segment of DNA through the other by using a reversible, enzyme-bridged double strand break. The transient enzyme/DNA adduct is mediated by a phosphodiester bond between the active-site tyrosine and a backbone phosphate of DNA. The opening and closing of the DNA gate, a critical step for strand passage during the catalytic cycle, is coupled to this cleavage/religation. We designed a unique oligonucleotide substrate with a pair of fluorophores straddling the topoisomerase II cleavage site, allowing the use of FRET to monitor the opening of the DNA gate. The DNA substrate undergoes an enzyme-mediated transition between a closed and open state in the presence of ATP, similar to the overall topoisomerase II catalyzed reaction. Single-molecule fluorescence microscopy measurements demonstrate that the transition has comparable rate constants for both the opening and closing reaction during steady-state ATP hydrolysis, with an apparent equilibrium constant near unity. In the presence of AMPPNP, a reduction in FRET occurs, suggesting an opening or partial opening of the DNA gate. However, the single-molecule experiments indicate that the open and closed states do not interconvert at a measurable rate.


Subject(s)
DNA Topoisomerases, Type II/metabolism , DNA/chemistry , DNA/metabolism , Drosophila melanogaster/enzymology , Eukaryotic Cells/enzymology , Nucleic Acid Conformation , Animals , Base Sequence , DNA/genetics , Fluorescence Resonance Energy Transfer , Fluorescent Dyes , Kinetics , Molecular Sequence Data , Substrate Specificity
10.
Chem Rev ; 106(8): 3080-94, 2006 Aug.
Article in English | MEDLINE | ID: mdl-16895319
11.
Biochemistry ; 45(26): 7990-7, 2006 Jul 04.
Article in English | MEDLINE | ID: mdl-16800624

ABSTRACT

In T4 bacteriophage, the DNA polymerase holoenzyme is responsible for accurate and processive DNA synthesis. The holoenzyme consists of DNA polymerase gp43 and clamp protein gp45. To form a productive holoenzyme complex, clamp loader protein gp44/62 is required for the loading of gp45, along with MgATP, and also for the subsequent binding of polymerase to the loaded clamp. Recently published evidence suggests that holoenzyme assembly in the T4 replisome may take place via more than one pathway [Zhuang, Z., Berdis, A. J., and Benkovic, S. J. (2006) Biochemistry 45, 7976-7989]. To demonstrate unequivocally whether there are multiple pathways leading to the formation of a productive holoenzyme, single-molecule fluorescence microscopy has been used to study the potential clamp loading and holoenzyme assembly pathways on a single-molecule DNA substrate. The results obtained reveal four pathways that foster the formation of a functional holoenzyme on DNA: (1) clamp loader-clamp complex binding to DNA followed by polymerase, (2) clamp loader binding to DNA followed by clamp and then polymerase, (3) clamp binding to DNA followed by clamp loader and then polymerase, and (4) polymerase binding to DNA followed by the clamp loader-clamp complex. In all cases, MgATP is required. The possible physiological significance of the various assembly pathways is discussed in the context of replication initiation and lagging strand synthesis during various stages of T4 phage replication.


Subject(s)
Bacteriophage T4/enzymology , DNA, Viral/metabolism , DNA-Directed DNA Polymerase/genetics , DNA-Directed DNA Polymerase/metabolism , Bacteriophage T4/genetics , Base Sequence , DNA, Viral/chemistry , DNA, Viral/genetics , Kinetics , Molecular Sequence Data , Protein Binding , Trans-Activators/genetics , Trans-Activators/metabolism , Viral Proteins/genetics , Viral Proteins/metabolism
12.
Biochemistry ; 44(51): 16835-43, 2005 Dec 27.
Article in English | MEDLINE | ID: mdl-16363797

ABSTRACT

Ensemble kinetics and single-molecule fluorescence microscopy were used to study conformational transitions associated with enzyme catalysis by dihydrofolate reductase (DHFR). The active site loop of DHFR was labeled with a fluorescence quencher, QSY35, at amino acid position 17, and the fluorescent probe, Alexa555, at amino acid 37, by introducing cysteines at these sites with site-specific mutagenesis. The distance between the probes was such that approximately 50% fluorescence resonance energy transfer (FRET) occurred. The double-labeled enzyme retained essentially full catalytic activity, and stopped-flow studies of both the forward and reverse reactions revealed that the distance between probes increased prior to hydride transfer. A fluctuation in fluorescence intensity of single molecules of DHFR was observed in an equilibrium mixture of substrates but not in their absence. Ensemble rate constants were derived from the distributions of lifetimes observed and attributed to a reversible conformational change. Studies were carried out with both NADPH and NADPD as substrates, with no measurable isotope effect. Similar studies with a G121V mutant DHFR resulted in smaller rate constants. This mutant DHFR has reduced catalytic activity, so that the collective data for the conformational change suggest that the conformational change being observed is associated with catalysis and probably represents a conformational change prior to hydride transfer. If the change in fluorescence is attributed to a change in FRET, the distance change associated with the conformational change is approximately 1-2 A. These results are correlated with other measurements related to conformation coupled catalysis.


Subject(s)
Escherichia coli/enzymology , Tetrahydrofolate Dehydrogenase/chemistry , Algorithms , Biotinylation , Catalysis , Cysteine/genetics , Fluorescence Resonance Energy Transfer , Fluorescent Dyes/chemistry , Folic Acid/analogs & derivatives , Folic Acid/chemistry , Hydrogen-Ion Concentration , Kinetics , Least-Squares Analysis , Models, Chemical , Mutagenesis, Site-Directed , Mutation/genetics , NADP/chemistry , Protein Conformation , Recombinant Proteins/chemistry , Recombinant Proteins/genetics , Recombinant Proteins/metabolism , Spectrometry, Fluorescence , Statistical Distributions , Tetrahydrofolate Dehydrogenase/genetics , Tetrahydrofolate Dehydrogenase/metabolism
13.
Biochemistry ; 44(37): 12420-33, 2005 Sep 20.
Article in English | MEDLINE | ID: mdl-16156655

ABSTRACT

R67 dihydrofolate reductase (DHFR) is a novel bacterial protein that possesses 222 symmetry and a single active site pore. Although the 222 symmetry implies that four symmetry-related binding sites must exist for each substrate as well as for each cofactor, various studies indicate only two molecules bind. Three possible combinations include two dihydrofolate molecules, two NADPH molecules, or one substrate plus one cofactor. The latter is the productive ternary complex. To explore the role of various ligand substituents during binding, numerous analogues, inhibitors, and fragments of NADPH and/or folate were used in both isothermal titration calorimetry (ITC) and K(i) studies. Not surprisingly, as the length of the molecule is shortened, affinity is lost, indicating that ligand connectivity is important in binding. The observed enthalpy change in ITC measurements arises from all components involved in the binding process, including proton uptake. As a buffer dependence for binding of folate was observed, this likely correlates with perturbation of the bound N3 pK(a), such that a neutral pteridine ring is preferred for pairwise interaction with the protein. Of interest, there is no enthalpic signal for binding of folate fragments such as dihydrobiopterin where the p-aminobenzoylglutamate tail has been removed, pointing to the tail as providing most of the enthalpic signal. For binding of NADPH and its analogues, the nicotinamide carboxamide is quite important. Differences between binary (binding of two identical ligands) and ternary complex formation are observed, indicating interligand pairing preferences. For example, while aminopterin and methotrexate both form binary complexes, albeit weakly, neither readily forms ternary complexes with the cofactor. These observations suggest a role for the O4 atom of folate in a pairing preference with NADPH, which ultimately facilitates catalysis.


Subject(s)
Tetrahydrofolate Dehydrogenase/chemistry , Tetrahydrofolate Dehydrogenase/metabolism , Calorimetry , Escherichia coli/enzymology , Escherichia coli/genetics , Folic Acid/metabolism , Folic Acid Antagonists/chemistry , Folic Acid Antagonists/pharmacology , Kinetics , Ligands , Models, Molecular , NADP/metabolism , Protein Conformation , Recombinant Proteins/chemistry , Recombinant Proteins/metabolism , Thermodynamics
14.
J Biol Chem ; 279(45): 46995-7002, 2004 Nov 05.
Article in English | MEDLINE | ID: mdl-15333636

ABSTRACT

R67 dihydrofolate reductase (R67 DHFR) is a novel protein encoded by an R-plasmid that confers resistance to the antibiotic, trimethoprim. This homotetrameric enzyme possesses 222 symmetry, which imposes numerous constraints on the single active site pore, including a "one-site-fits-both" strategy for binding its ligands, dihydrofolate (DHF) and NADPH. Previous studies uncovered salt effects on binding and catalysis (Hicks, S. N., Smiley, R. D., Hamilton, J. B., and Howell, E. E. (2003) Biochemistry 42, 10569-10578), however the one or more residues that participate in ionic contacts with the negatively charged tail of DHF as well as the phosphate groups in NADPH were not identified. Several studies predict that Lys-32 residues were involved, however mutations at this residue destabilize the R67 DHFR homotetramer. To study the role of Lys-32 in binding and catalysis, asymmetric K32M mutations have been utilized. To create asymmetry, individual mutations were added to a tandem array of four in-frame gene copies. These studies show one K32M mutation is tolerated quite well, whereas addition of two mutations has variable effects. Two double mutants, K32M:1+2 and K32M: 1+4, which place the mutations on opposite sides of the pore, reduce kcat. However a third double mutant, K32M: 1+3, that places two mutations on the same half pore, enhances kcat 4- to 5-fold compared with the parent enzyme, albeit at the expense of weaker binding of ligands. Because the kcat/Km values for this double mutant series are similar, these mutations appear to have uncovered some degree of non-productive binding. This non-productive binding mode likely arises from formation of an ionic interaction that must be broken to allow access to the transition state. The K32M:1+3 mutant data suggest this interaction is an ionic interaction between Lys-32 and the charged tail of dihydrofolate. This unusual catalytic scenario arises from the 222 symmetry imposed on the single active site pore.


Subject(s)
Lysine/chemistry , Mutation , Tetrahydrofolate Dehydrogenase/biosynthesis , Tetrahydrofolate Dehydrogenase/chemistry , Binding Sites , Catalysis , Circular Dichroism , Dimerization , Dose-Response Relationship, Drug , Escherichia coli/enzymology , Escherichia coli/metabolism , Hydrogen-Ion Concentration , Ions , Kinetics , Ligands , Models, Biological , Models, Molecular , NADP/chemistry , Protein Binding , Protein Conformation , Salts/pharmacology , Spectrometry, Fluorescence , Temperature
15.
J Biol Chem ; 279(45): 47003-9, 2004 Nov 05.
Article in English | MEDLINE | ID: mdl-15333637

ABSTRACT

R67 dihydrofolate reductase (R67 DHFR) catalyzes the transfer of a hydride ion from NADPH to dihydrofolate, generating tetrahydrofolate. The homotetrameric enzyme provides a unique environment for catalysis as both ligands bind within a single active site pore possessing 222 symmetry. Mutation of one active site residue results in concurrent mutation of three additional symmetry-related residues, causing large effects on binding of both ligands as well as catalysis. For example, mutation of symmetry-related tyrosine 69 residues to phenylalanine (Y69F), results in large increases in Km values for both ligands and a 2-fold rise in the kcat value for the reaction (Strader, M. B., Smiley, R. D., Stinnett, L. G., VerBerkmoes, N. C., and Howell, E. E. (2001) Biochemistry 40, 11344-11352). To understand the interactions between specific Tyr-69 residues and each ligand, asymmetric Y69F mutants were generated that contain one to four Y69F mutations. A general trend observed from isothermal titration calorimetry and steady-state kinetic studies of these asymmetric mutants is that increasing the number of Y69F mutations results in an increase in the Kd and Km values. In addition, a comparison of steady-state kinetic values suggests that two Tyr-69 residues in one half of the active site pore are necessary for NADPH to exhibit a wild-type Km value. A tyrosine 69 to leucine mutant was also generated to approach the type(s) of interaction(s) occurring between Tyr-69 residues and the ligands. These studies suggest that the hydroxyl group of Tyr-69 is important for interactions with NADPH, whereas both the hydroxyl group and hydrophobic ring atoms of the Tyr-69 residues are necessary for proper interactions with dihydrofolate.


Subject(s)
Tetrahydrofolate Dehydrogenase/chemistry , Tyrosine/chemistry , Binding Sites , Calorimetry , Catalysis , Dimerization , Escherichia coli/enzymology , Hydrogen-Ion Concentration , Ions , Kinetics , Leucine/chemistry , Ligands , Models, Chemical , Models, Molecular , Mutagenesis, Site-Directed , Mutation , NADP/chemistry , Protein Binding , Protein Structure, Tertiary , Thermodynamics
16.
Biochemistry ; 43(23): 7403-12, 2004 Jun 15.
Article in English | MEDLINE | ID: mdl-15182183

ABSTRACT

R67 dihydrofolate reductase (DHFR) is a novel protein that possesses 222 symmetry. A single active site pore traverses the length of the homotetramer. Although the 222 symmetry implies that four symmetry-related binding sites should exist for each substrate as well as each cofactor, isothermal titration calorimetry (ITC) studies indicate only two molecules bind. Three possible combinations include two dihydrofolate molecules, two NADPH molecules, or one substrate with one cofactor. The latter is the productive ternary complex. To evaluate the roles of A36, Y46, T51, G64, and V66 residues in binding and catalysis, a site-directed mutagenesis approach was employed. One mutation per gene produces four mutations per active site pore, which often result in large cumulative effects. Conservative mutations at these positions either eliminate the ability of the gene to confer trimethoprim resistance or have no effect on catalysis. This result, in conjunction with previous mutagenesis studies on K32, K33, S65, Q67, I68, and Y69 [Strader, M. B., et al. (2001) Biochemistry 40, 11344-11352; Hicks, S. N., et al. (2003) Biochemistry 42, 10569-10578; Park, H., et al. (1997) Protein Eng. 10, 1415-1424], allows mapping of the active site surface. Residues for which conservative mutations have large effects on binding and catalysis include K32, Q67, I68, and Y69. These residues form a stripe that establishes the ligand binding surface. Residues that accommodate conservative mutations that do not greatly affect catalysis include K33, Y46, T51, S65, and V66. Isothermal titration calorimetry studies were also conducted on many of the mutants described above to determine the enthalpy of folate binding to the R67 DHFR.NADPH complex. A linear correlation between this DeltaH value and log k(cat)/K(m) is observed. Since structural tightness appears to be correlated with the exothermicity of the binding interaction, this leads to the hypothesis that enthalpy-driven formation of the ternary complex in these R67 DHFR variants plays a strong role in catalysis. Use of the alternate cofactor, NADH, extends this correlation, indicating preorganization of the ternary complex determines the efficiency of the reaction. This hypothesis is consistent with data suggesting R67 DHFR uses an endo transition state (where the nicotinamide ring of cofactor overlaps the more bulky side of the substrate's pteridine ring).


Subject(s)
Tetrahydrofolate Dehydrogenase/chemistry , Tetrahydrofolate Dehydrogenase/metabolism , Amino Acids/chemistry , Amino Acids/genetics , Amino Acids/metabolism , Binding Sites , Catalysis , Dimerization , Hydrogen-Ion Concentration , Kinetics , Models, Molecular , Mutagenesis, Site-Directed/genetics , Mutation/genetics , NADP/chemistry , NADP/metabolism , Protein Structure, Quaternary , Tetrahydrofolate Dehydrogenase/genetics , Thermodynamics
17.
Biochemistry ; 42(36): 10569-78, 2003 Sep 16.
Article in English | MEDLINE | ID: mdl-12962480

ABSTRACT

R67 dihydrofolate reductase (DHFR), which catalyzes the NADPH dependent reduction of dihydrofolate to tetrahydrofolate, belongs to a type II family of R-plasmid encoded DHFRs that confer resistance to the antibacterial drug trimethoprim. Crystal structure data reveals this enzyme is a homotetramer that possesses a single active site pore. Only two charged residues in each monomer are located near the pore, K32 and K33. Site-directed mutants were constructed to probe the role of these residues in ligand binding and/or catalysis. As a result of the 222 symmetry of this enzyme, mutagenesis of one residue results in modification at four related sites. All mutants at K32 affected the quaternary structure, producing an inactive dimer. The K33M mutant shows only a 2-4-fold effect on K(m) values. Salt effects on ligand binding and catalysis for K33M and wildtype R67 DHFRs were investigated to determine if these lysines are involved in forming ionic interactions with the negatively charged substrates, dihydrofolate (overall charge of -2) and NADPH (overall charge of -3). Binding studies indicate that two ionic interactions occur between NADPH and R67 DHFR. In contrast, the binding of folate, a poor substrate, to R67 DHFR.NADPH appears weak as a titration in enthalpy is lost at low ionic strength. Steady-state kinetic studies for both wild type (wt) and K33M R67 DHFRs also support a strong electrostatic interaction between NADPH and the enzyme. Interestingly, quantitation of the observed salt effects by measuring the slopes of the log of ionic strength versus the log of k(cat)/K(m) plots indicates that only one ionic interaction is involved in forming the transition state. These data support a model where two ionic interactions are formed between NADPH and symmetry related K32 residues in the ground state. To reach the transition state, an ionic interaction between K32 and the pyrophosphate bridge is broken. This unusual scenario likely arises from the constraints imposed by the 222 symmetry of the enzyme.


Subject(s)
R Factors/chemistry , R Factors/metabolism , Tetrahydrofolate Dehydrogenase/chemistry , Tetrahydrofolate Dehydrogenase/metabolism , Calorimetry/methods , Catalysis , Dimerization , Escherichia coli Proteins/chemistry , Escherichia coli Proteins/genetics , Escherichia coli Proteins/metabolism , Folic Acid/analogs & derivatives , Folic Acid/metabolism , Hydrogen-Ion Concentration , Kinetics , Ligands , Models, Molecular , Mutagenesis, Site-Directed , NADP/metabolism , Osmolar Concentration , Protein Binding , R Factors/genetics , Sodium/chemistry , Spectrometry, Fluorescence/methods , Tetrahydrofolate Dehydrogenase/genetics , Thermodynamics
18.
Biochemistry ; 41(52): 15664-75, 2002 Dec 31.
Article in English | MEDLINE | ID: mdl-12501195

ABSTRACT

R67 dihydrofolate reductase (DHFR) is an enzyme, encoded by an R-plasmid, that confers resistance to the antibacterial agent trimethoprim. This homotetramer possesses a single active site pore and exact 222 symmetry. The symmetry imposes constraints on the ability of the enzyme to optimize binding of the substrate, dihydrofolate (DHF), and the cofactor, NADPH, resulting in a "one site fits both ligands" approach. This approach allows formation of either a NADPH.NADPH, dihydrofolate.dihydrofolate, or NADPH.dihydrofolate complex. The first two complexes are nonproductive, while the third is the productive catalytic species. To break the symmetry of the active site, a tandem array of four R67 DHFR genes has been linked in frame, allowing individual manipulation of each gene copy. Various numbers and combinations of asymmetric Q67H mutations have been engineered into the tandem gene array. The Q67H mutation was chosen for investigation as it was previously found to tighten binding to both dihydrofolate and NADPH by approximately 100-fold in homotetrameric R67 DHFR [Park, H., Bradrick, T. D., and Howell, E. E. (1997) Protein Eng. 10, 1415-1424]. Nonadditive effects on ligand binding are observed when one to four mutations are inserted, indicating either conformational changes in the protein or different cooperativity patterns in the ligand-ligand interactions. From steady state kinetics, addition of Q67H mutations does not drastically affect formation of the NADPH.dihydrofolate complex; however, a large energy difference between the productive and nonproductive complexes is no longer maintained. A role for Q67 in discriminating between these various states is proposed. Since theories of protein evolution suggest gene duplication followed by accumulation of mutations can lead to divergence of activity, this study is a first step toward asking if introduction of asymmetric mutations in the quadrupled R67 DHFR gene can lead to optimization of ligand binding sites.


Subject(s)
Mutagenesis, Site-Directed , Tetrahydrofolate Dehydrogenase/chemistry , Tetrahydrofolate Dehydrogenase/genetics , Amino Acid Substitution/genetics , Binding Sites/genetics , Circular Dichroism , Gene Dosage , Genes, Bacterial/genetics , Genes, Synthetic/genetics , Glutamine/genetics , Histidine/genetics , Kinetics , Molecular Mimicry , NADP/chemistry , R Factors/genetics , Tandem Repeat Sequences/genetics
19.
Anal Biochem ; 305(1): 68-81, 2002 Jun 01.
Article in English | MEDLINE | ID: mdl-12018947

ABSTRACT

Mass spectrometry is currently the method of choice for the analysis of recombinant protein expression products. By combining proteolytic digestion with peptide mapping and tandem mass spectrometry techniques, verification of site-directed mutagenesis products can be obtained. The proteolytic digestion step converts a purified recombinant protein into a mixture that must be reseparated, thus greatly increasing the analysis time associated with the confirmation of site-directed mutagenesis products. Ion/ion reaction chemistry combined with quadrupole ion trap mass spectrometry provides a fast and efficient way to analyze intact proteins for the correct site-directed mutagenesis products, without heavy reliance on the proteolytic digestion step. Analysis of a series of protein variants (I68M, I68Q, Y69F, and Q67Y) from plasmid-encoded R67 dihydrofolate reductase using ion/ion reaction chemistry confirmed the presence of the correct site-directed mutagenesis products. For the I68M mutant, ion/ion separations detected the presence of extensive degradation from the N-terminal end of the protein. In the case of the Q67Y mutant, a mixture of Q67Y and Q67C species was detected by employing tandem mass spectrometry combined with ion/ion reactions. The ion/ion reaction technique was also performed on a partially purified lysate of the Q67Y/C mixture and successfully screened for the presence of both components in a complex mixture. The ion/ion reaction approach achieved the same results as the proteolytic-digestion-based methodology in a much shorter analysis time.


Subject(s)
Mutagenesis, Site-Directed/genetics , Peptide Fragments/analysis , Recombinant Proteins/analysis , Tetrahydrofolate Dehydrogenase/analysis , Amino Acid Sequence , Amino Acids/genetics , Binding Sites , Chromatography/methods , Electrophoresis, Polyacrylamide Gel , Escherichia coli/enzymology , Escherichia coli/genetics , Mass Spectrometry/methods , Molecular Sequence Data , Mutation/genetics , Peptide Fragments/chemistry , Peptide Mapping , Plasmids/genetics , Recombinant Proteins/chemistry , Recombinant Proteins/genetics , Serine Endopeptidases/chemistry , Serine Endopeptidases/metabolism , Spectrometry, Mass, Electrospray Ionization , Spectrometry, Mass, Matrix-Assisted Laser Desorption-Ionization , Tetrahydrofolate Dehydrogenase/genetics , Tetrahydrofolate Dehydrogenase/metabolism , Time Factors
SELECTION OF CITATIONS
SEARCH DETAIL
...